• Keine Ergebnisse gefunden

The role of aggregation for the dissolution of diatom frustules Uta Passow

N/A
N/A
Protected

Academic year: 2022

Aktie "The role of aggregation for the dissolution of diatom frustules Uta Passow"

Copied!
9
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

The role of aggregation for the dissolution of diatom frustules

Uta Passow

a;

, Anja Engel

a

, Helle Ploug

b

a Alfred-Wegener-Institut fu«r Polar und Meeresforschung, Bremerhaven, Germany

b Max-Planck Institut fu«r Marine Mikrobiologie, Bremen, Germany

Received 15 November 2002 ; received in revised form 26 June 2003 ; accepted 7 July 2003 First published online 27 August 2003

Abstract

Observations that the majority of silica dissolution occurs within the upper 200 m of the ocean, and that sedimentation rates of diatom frustules generally do not decrease significantly with depth, suggested reduced dissolution rates of diatoms embedded within sinking aggregates. To investigate this hypothesis, silica dissolution rates of aggregated diatom cells were compared to those of dispersed cells during conditions mimicking sedimentation below the euphotic zone. Changes in the concentrations of biogenic silica, silicic acid, cell numbers, chlorophyllaand transparent exopolymer particles (TEP) were monitored within aggregates and in the surrounding seawater (SSW) during two 42-day experiments. Whereas the concentration of dispersed diatoms decreased over the course of the experiment, the amount of aggregated cells remained roughly constant after an initial increase. Initially only 6% of cells were aggregated and at the end of the experiment more than 60% of cells were enclosed within aggregates. These data imply lower dissolution rates for aggregated cells.

However, fluxes of silica between the different pools could not be constrained reliably enough to unequivocally prove reduced dissolution for aggregated cells.

9 2003 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.

Keywords : Diatom frustule ; Silica dissolution ; Aggregation

1. Introduction

The rate of silica dissolution from diatom frustules de- termines the fraction of silica recycled within the euphotic zone where it is available for new growth, as well as the fraction of biogenic silica which is recycled at depth or buried forming diatomaceous ooze. Even though the sol- ubility of biogenic silica (BSiO2) in seawater is very high, s1000WM at 15‡C[1]and the ocean is undersaturated in silicic acid (upper ocean : 0^50 WM, deep ocean : 6180 WM[2]), dissolution rates of diatom frustules are generally low. Diatom frustules are protected from immediate dis- solution by an organic coating consisting of three layers, a polysaccharide, a lipid and a protein layer [3]. High dis- solution rates of diatom frustules occur when bacterial degradation removes this protective organic matrix[4,5].

Diatoms settle to the deep ocean predominantly as ag- gregates[6]. Measurements within aggregates indicate that

the microenvironment of aggregates di¡ers considerably from environmental conditions in the surrounding water [7^10]. Consequently, the rate of biological and chemical processes within aggregates may di¡er fundamentally from those in the surrounding seawater (SSW). Dissolution rates of diatom frustules, for example, may be reduced or elevated within aggregates compared to those of freely suspended cells. High bacterial activity within aggregates may increase average dissolution rates due to the removal of the protective organic matrix. However, if the exchange rate between the pore water of aggregates and the SSW is low, the accumulation of silicic acid within aggregates may decrease dissolution rates of diatom frustules, because once the organic coating is removed, dissolution rate is a function of the concentration gradient between frustules and the adjacent water. Preliminary experimental evidence suggested relatively low dissolution rates (1.6% day31) when phytodetritus was aggregated compared to non-ag- gregated material [11], but a di¡erent study did not ob- serve di¡erences in dissolution between clumped and free cells [4]. Dissolution rates of frustules within intact fecal pellets are reduced[12], suggesting that frustules packed in marine snow may also be protected from dissolution. A

0168-6496 / 03 / $22.00 9 2003 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.

* Corresponding author. Tel. : +49 (471) 4831 1450 ; Fax : +49 (471) 4831 1425.

E-mail address :upassow@awi-bremerhaven.de(U. Passow).

www.fems-microbiology.org

(2)

minimal decrease in sedimentation rates of diatoms with depth, also imply that dissolution of frustules is insignif- icant during marine snow-mediated sinking[13].

In this study we investigated the dissolution of diatom cells during conditions mimicking their sedimentation as aggregated or dispersed cells. We hypothesized that the average dissolution rates of aggregated cells would be ap- preciably lower than that of freely suspended cells. Specif- ically, we monitored silica dissolution of the diatomTha- lassiosira weiss£ogii, sampling aggregated and freely dispersed cells separately. Concentrations of bacteria and transparent exopolymer particles (TEP) were also moni- tored during the experiments.

2. Materials and methods

2.1. Experimental approach and set-up

The experiments were set-up to compare silica dissolu- tion of living T. weiss£ogii cells within aggregates with those of dispersed cells in the SSW during conditions imi- tating continuous sinking in the dark (Fig. 1). A multi- bottle approach, where replicate bottles were sacri¢ced at each time point, was chosen, as it is nearly impossible to subsample from a population of aggregates, because of their high variability (e.g. cell numbers of di¡erent aggre- gates of similar size may vary widely). Aggregated and dispersed cells were incubated together and separated pri- or to analysis rather than before incubation, to guarantee

identical start-up conditions in each bottle. A separation of aggregates and dispersed cells before incubation is pre- carious, because aggregates easily break or get damaged during handling and because it is impossible to prepare enough aggregates of the same size and cell concentration to generate a set of replicate bottles of aggregates.

Before the experiment, the solitary, centric central dia- tomT. weiss£ogiiwas grown in two batch cultures at 15‡C with a light £ux of 100 Wmol m32 s31 in a 12h :12h light :dark cycle. F/2media based on seawater was used [14], except that nutrient concentrations were reduced to 30% to avoid arti¢cially high cell concentrations. Nutrient concentrations were kept non-limiting in one culture, whereas silicic acid concentration was run to depletion in the second culture to ensure a low background concentra- tion of silicic acid at the start of the experiment. Two identical experiments, experiments A and B, were con- ducted. Both experiments were designed to last for 42 days with nine sampling days spread evenly over the in- cubation period. Because three replicate bottles were to be sacri¢ced on each sampling day, a total of 27 gas-perme- able bottles (LDPE, Nalgene) with a volume of 116 ml each were prepared at the beginning of each experiment.

Each bottle was ¢lled with a mixture of growing diatoms from the ¢rst culture (nitrate : 184 WM, phosphate : 2.4 WM, silicic acid : 9.6 WM) and diatom-free ¢ltrate gener- ated from the second culture (nitrate : 185WM, phosphate : 2.9WM, silicic acid : 0.7WM) to achieve an end concentra- tion of about 3000 cells ml31. Each bottle was ¢lled bub- ble-free, by closing the bottle while submerged. Bottles

Fig. 1. Experimental set-up : Two identical experiments lasting 42days were conducted in the dark. Diatoms grown in silicic acid reduced media were initially incubated on a rolling table for 48 h to promote aggregate formation and then incubated on a plankton (Ferris) wheel to simulate continuous sinking. At each sampling date three replicate bottles were removed and the aggregates separated from the SSW. Both fractions were then analyzed sep- arately.

(3)

and lids were handled with sterilized tongs while sub- merged in the diatom suspension. Bubbles would create turbulence and shear in bottles rotating on a Ferris wheel, increasing collisions between cells, aggregates and bottle walls, rather than mimicking sinking. The experiments were conducted in the dark at a temperature of 15‡C.

During the ¢rst 48 h bottles were incubated on a roller table to promote aggregation [15]. When aggregates were visible, the bottles were transferred carefully to a Ferris wheel (diameter = 0.9 m) and rotated continuously end over end at about 1 rpm to simulate continuous sinking in a low-shear environment.

Visible aggregates (v1 mm) and the SSW without visi- ble aggregates were analyzed separately. On each sampling day, aggregates were isolated with a 10 ml syringe equipped with a 1 mm diameter needle. The total volume of the aggregate slurry for each bottle was recorded to the nearest 0.1 ml, and 35x NaCl solution was added to bring the total volume to 100 ml. The remaining volume of SSW was also documented for each bottle.

2.2. Analysis

On each sampling day, subsamples for the determina- tion of the following variables were taken from both frac- tions (the aggregate slurry and the SSW fraction) of each of three replicate bottles. Nutrients (NO3, PO4 and Si(OH)4) were analyzed with an autoanalyzer[16]. Biogen- ic silica was measured from 50^60 ml sample ¢ltered onto 0.4 Wm polycarbonate ¢lters and analyzed according to Korole¡ [16]. Chlorophyll a was determined in a Turner

£uorometer [17] from two replicate GF/F ¢lters onto which 4^8 ml of sample were ¢ltered. Phytoplankton abundance was determined from 20 ml samples, which were preserved with 4 ml of formalin (20%). Phytoplank- ton cells were counted using an inverted microscope (Zeiss) at magni¢cations of 250U and 400U. TEP were analyzed colorimetrically from 20 ml[18]and microscopi- cally from 10 ml ¢ltered onto 0.4 Wm membrane ¢lters (Poretics) with two replicates each [19]. Two replicate black 0.2 Wm membrane ¢lters (Poretics) were prepared from 2^4 ml sample, ¢xed with 0.2Wm pre-¢ltered forma- lin (40%) and stained with 0.5 ml of pre-¢ltered (0.2Wm) 4,6-diamino-2-phenylindole (DAPI) for the enumeration of bacteria. Stained ¢lters were mounted onto glass slides, covered with immersion oil and a glass cover and frozen at 320‡C. Slides were then transferred to a £uorescent light microscope (Zeiss, 340 nm light source) and screened by a Panasonic color video camera on Super VHS at a magni-

¢cation of 1000U. Fifty frames were chosen in a cross- section and digitized on a Macintosh PPC with an optical resolution of 0.028 Wm2 per pixel. Bacteria were enumer- ated and sized semi-automatically by the image analysis program NIH-Image 6.1 ppc, a public domain program developed at the US National Institute of Health.

Whereas it was straightforward to calculate the concen-

trations of the particulate and dissolved constituents in the SSW, as well as the concentrations of particulates enclosed within aggregates from measured amounts and volumes, calculations of the concentrations of solutes in the pore water of aggregates are prone to high errors. The determi- nation of the pore water volume necessitates a set of as- sumptions regarding the porosity [20^22] and fractal na- ture of aggregates [23^26], the size of the boundary layer and plume of aggregates [9,27,28], the £ow ¢elds around or through aggregates [20,29]and the molecular di¡usion coe⁄cient of silicic acid out of aggregates [10]. Past esti- mates of nutrient concentrations in pore water of aggre- gates may be too high, if they are based on calculations ignoring the newly hypothesized plume surrounding aggre- gates. The volume of water characterized by elevated con- centrations includes both the pore water of the aggregate and the water in the plume. Several of the estimates needed to calculate pore water concentrations are cur- rently not well constrained and discussed very controver- sially. Here we thus only present the data of solutes in pore water as absolute amounts, not concentrations.

3. Results and discussion

We analyzed data of both experiments, but because ini- tial concentrations of dissolved and particulate matter were not signi¢cantly di¡erent and because all results of

Fig. 2. Overall development of nutrient ((a) nitrate, (b) phosphate and (c) silicic acid) concentrations in bottles (aggregate and SSW fractions combined) during experiment A. Averages and standard deviation of the three replicate bottles are depicted.

(4)

experiments A and B showed the same trend, only data from experiment A are presented.

3.1. General development of dissolved and particulate constituents

To sketch the overall development during the experi- ment, averages of the total concentrations of nutrients (Fig. 2) and particles (Fig. 3) are depicted from the three replicate bottles during the 42-day period. Nitrate concen- tration was high throughout and increased slightly during the ¢rst 20 days of the experiment (Fig. 2a). Initially, phosphate concentration was low (0.4 WM) and after a lag period of 12days, increased exponentially to a ¢nal value of 6.4WM (Fig. 2b). At the beginning of the experi- ment, the concentration of silicic acid was low (5.6 WM), but increased linearly from the beginning of the experi- ment to 34 WM indicating rapid dissolution of biogenic silica from the start (Fig. 2c).

In contrast to chlorophyll a, which began decreasing during the ¢rst days of the study, the diatom abundance remained constant during the ¢rst 2weeks but decreased by roughly two thirds thereafter (Fig. 3a,b). Biogenic silica began decreasing without delay (Fig. 3d), mirroring the

increase in silicic acid concentration (Fig. 4). The total

£ux of silica from the biogenic into the dissolved pool was equal to 5 Wmol per bottle during the 42-day experi- ment. Whereas dissolution of silica frustules and the de- cline of chlorophyll a began within 1 or 2days after the onset of darkness, the loss of cells and the degradation of phosphor were delayed for almost 2weeks after the onset of darkness, suggesting a temporal decoupling between the degradation of cells and silica dissolution.

Initial concentrations of TEP were high (8000 Wg xan- than equivalent l31) and decreased by one half during the

¢rst 20 days of the experiment, but remained almost con- stant thereafter (Fig. 3e). TEP are a chemically heteroge- neous group of particles and whereas one fraction was degraded on timescales of days, degradation of the re- maining fraction may have been appreciably lower. Alter- natively enhanced production of TEP by bacteria may have compensated loss by degradation during the second half of the experiment. Total bacterial concentration ini- tially decreased and then £uctuated around 2U109 cells l31 with a slight decreasing tendency (Fig. 3c).

3.2. Developments of dissolved and particulate constituents within the di¡erent compartments

The dynamics of diatom and bacteria numbers, as well as of chlorophylla, biogenic silica and TEP concentrations in the SSW fraction re£ected predominantly the decrease observed in the SSW plus aggregate fraction presented above, except that the decreases were more pronounced in the SSW fractions compared to the combined fractions.

The amount of chlorophyll a, biogenic silica, diatoms and TEP collected in the aggregate slurry increased during the ¢rst 2weeks of the study (Fig. 5). During the following 10 days, variability between replicate bottles was high, but

Fig. 3. Overall development of the (a) chlorophyll a, (b) diatoms, (c) bacteria, (d) BSi and (e) TEP concentrations in bottles (aggregate and SSW fractions combined) during experiment A. Averages and standard deviation of the three replicate bottles are depicted.

Fig. 4. Overall development of dissolved and biogenic silica concentra- tions in bottles (aggregate and SSW fractions combined) during experi- ment A. Averages and standard deviation of the three replicate bottles are given. Note the non-linear timescale.

(5)

no net change was visible. After day 25 the concentrations of chlorophyll a, biogenic silica, cell number and TEP in the aggregate slurry were high, but decreased during the remainder of the study. Less than 20% of TEP or diatoms were incorporated in the visible aggregates before the pe- riod of high variability, but after day 25 more than 50% of these particles were sampled in the aggregate slurry (Fig.

6). No net change in bacterial numbers within aggregate slurries was observed during the initial 15 days (Fig. 5c).

Bacteria concentrations within aggregates (2.2 S 1.0U107 cells ml31) were low and on average only one order of magnitude higher than in the SSW, where abundance was comparable to natural seawater with 1.2S 3.7U106 cells ml31.

Qualitatively, the aggregated material did not di¡er in a statistically signi¢cant way (t-test,K= 0.05) from the freely suspended material, as indicated by the ratios between chlorophyll a, biogenic silica or TEP and cell numbers, with one exception. During the initial 15 days of the ex- periment the chlorophyll aconcentration per cell was sig- ni¢cantly lower in aggregated compared to free material, suggesting either preferential loss of chlorophyll a within aggregates or preferential aggregation of chlorophyll a-poor cells. On each day, the biogenic silica content per

cell was higher in aggregated cells compared to free cells, but overall this di¡erence between aggregated and free cells was not statistically signi¢cant, because of the high day-to-day variability. The ratio between bacteria and di- atoms increased in SSW and £uctuated without trend in aggregates, but the average ratio was similar in both frac- tions.

3.3. Flux of silica between the di¡erent pools

Silica was measured in all four compartments ; as bio- genic (particulate) silica in the SSW and in aggregates and as silicic acid both in the SSW and within pore water of aggregates. The net decrease of biogenic silica in the SSW

Fig. 5. Overall development of the amount of (a) chlorophylla, (b) dia- toms, (c) bacteria, (d) BSi and (e) TEP and sampled in aggregate slur- ries during experiment A. Averages and standard deviation of the three replicate bottles are depicted.

Fig. 6. Percentage of total amount of chlorophylla(¢lled squares), TEP (open triangles) and diatoms (open circles) contained in aggregates dur- ing experiment A.

Fig. 7. Silica budget, showing the contributions of the biogenic silica in aggregates (AGBSIO2), and the SSW (SSWBSIO2) as well as the silicic acid content in aggregates (AGSiðOHÞ4) and the SSW (SSWSiðOHÞ4).

(6)

was compensated by a net increase of silicic acid in the SSW and a smaller net increase in biogenic silica in aggre- gates (except during the last day) (Fig. 7). As the total amount of silicic acid in the pore water was small, the change in the silicic acid amount in the pore water was also very small. It is assumed that the formation of bio- genic silica due to growth was negligible during the study, as the experiments were conducted in the dark. Diatom uptake of silicic acid can continue for only a few hours to a day in the dark [30]. Net disaggregation was also as- sumed to be negligible, as aggregates appeared sturdy and did not break even during handling. Aggregation and dissolution of cells as well as the e¥ux of solutes out of aggregates are thus the major potential mechanisms responsible for the £ux of silica between the four pools (Fig. 8).

3.3.1. Dissolution

Concentrations of silicic acid within the dispersed cell fraction initially were 5WM and increased to 33WM at the end, covering a range well comparable to natural seawater and far below saturation concentrations which are s1000 WM [1,31]. Within the range covered during our experi- ment the overall dissolution rate of aggregated plus dis- persed cells was constant (r2= 0.97) at 0.079Wmol per day per bottle, indicating that dissolution during our experi- ment was independent of the biogenic silica concentration.

If the experiment were continued further, dissolution rate would decrease as the biogenic silica concentration de- creased. In the past, di¡erent models have been used to represent the kinetics of biogenic silica dissolution, de- pending on assumptions regarding the dissolution mecha- nism[32]. As silica dissolution remained in the linear sec- tion of the dissolution curve during our experiment (meaning it was independent of biogenic silica concentra- tion), dynamics of silica dissolution could be ¢tted using any of several models describing silica dissolution (for an overview of models see [32]). The simplest model, the

Nernst di¡usion approach assumes that dissolution is lim- ited by the rate of di¡usion through a boundary layer (r2= 0.97, df= 7). The surface reaction approach assumes that dissolution is limited by surface chemistry, rather than di¡usion (r2= 0.97, df= 7). The parabolic law ap- proach assumes that the reaction rate is limited by di¡u- sion, di¡erentiating between each dissolved species (r2= 0.99, df= 7) and the decreasing surface approach ad- ditionally assumes a decreasing surface area as dissolution progresses (r2= 0.94, df= 25).

Daily speci¢c dissolution rates of aggregated plus dis- persed cells normalized by the initial concentration of BSi decreased from 3.3 to 1.7% day31 (0.033 to 0.017 cells dissolved day31), with a linear average of 2.3 S 0.5%

day31 (0.023 S 0.005 cells dissolved day31) or a value of 2.6% day31 (0.026 cells dissolved day31) calculated from an exponential curve ¢t using the initial concentration of biogenic silica as a potential maximum (Fig. 9). More than 60% of biogenic silica dissolved during the 42-day experi- ment (Fig. 9). A dissolution rate of 40^70% of diatom frustules in 40^50 days was also found for several other diatom species, independent whether cells were alive or heat killed, whether they were fragmented or intact, as long as bacteria were present [4]. The main factor deter- mining the dissolution rate of diatom frustules is the pres- ence or absence of bacteria [4,5]. Dissolution rates in the absence of bacteria were found to be 2^20 times lower.

Bacteria-mediated speci¢c dissolution rates of fresh phyto- detritus of T. weiss£ogii ranged between 2.1 and 7.9%

day31 initially and between 2.0 and 3.2% day31 between day 2and 10 depending on the colonizing bacteria [5].

Contrary to this latter study no initial increase in bacterial

Fig. 8. Simple box model depicting the four silica pools and the major

£uxes between these pools. Agg. BSi, biogenic silica in aggregated par- ticles ; SSW BSi, biogenic silica of particles in the SSW ; pore w.

Si(OH)4, silicic acid in pore water of aggregates ; SSW Si(OH)4, silicic acid in the SSW.

Fig. 9. Dissolution dynamics of diatom frustules depicted as the increase in silicic acid concentration and as the % of biogenic silica solubilized.

Data points and standard deviations between replicate bottles are given.

Changes in the silicic acid dissolution ¢t the curve : Silicic acid = 42.36 (13e30:026t),r2= 0.99, df= 8.

(7)

numbers was observed in our study. Moreover, concentra- tions of bacteria in aggregates were low during our study compared to the much higher enrichment of bacteria often found in aggregates ; see compilation of data in[33].

A net decrease in biogenic silica was observed in SSW only, no net decrease of cells was observed in the aggre- gate slurry fraction, and after an initial increase the amount of silicic acid in pore water remained roughly constant, implying that no signi¢cant net dissolution oc- curred. Assuming that no exchange between the aggre- gated and the dispersed cell fractions took place, the dis- solution rate in SSW was 0.65 Wmol l31 day31 (linear, r2= 0.97) and the speci¢c dissolution rate normalized to daily biogenic silica concentrations increased during the experiment by an order of magnitude from 0.017 to 0.18 day31 due to the decreasing biogenic silica concentration.

In aggregates the dissolution rate was 0.01Wmol l31day31 (highly variable, no signi¢cant relationship) with a speci¢c dissolution rate ranging from 0.001 to 0.003 day31, again assuming that no exchange between the aggregated and the dispersed cell fractions took place. Possibly silicic acid concentration in pore water was high enough to in- hibit dissolution of aggregated cells. Or, diatom frustules within aggregates were protected chemically (by organic substances) reducing dissolution. Alternatively, dissolution of aggregated cells was higher than estimated above, but balanced by continuous aggregation. This scenario is sup- ported by the fact that the overall speci¢c dissolution rate didn’t decrease appreciably after 26 days, although 60%

of all diatoms were imbedded in aggregates from that day on. In this scenario, the concomitant production of silicic acid would have to be balanced by its e¥ux out of aggre- gates.

3.3.2. Aggregation

After the initial increase in the concentrations of dia- toms in the aggregate slurry fraction during the ¢rst 15 days of the experiment, no net changes were observed, indicating that no further aggregation occurred or that the increase of particles in aggregates due to aggregation was balanced by dissolution. Using a simple coagulation model we estimated a maximal aggregation rate to evalu- ate if the decrease in freely suspended cells during the experiment could theoretically be explained by their inclu- sion in aggregates, rather than by their dissolution. Initial calculations had suggested that the £ux from the dispersed fraction into the aggregate fraction was dominated by scavenging of dispersed cells by aggregates due to di¡er- ential settling, whereas aggregation of single cells with each other due to shear was comparably small. Assuming maximum aggregation, e.g. that cells were lost from the SSW by aggregation only, we calculated the maximal transfer of cells into the aggregates using the basic coag- ulation equations given by Jackson[34]under the assump- tion of two size classes ; single cells and large, visible ag- gregates:

C1;2¼KN1N2 f½Z=4 ðd1þd2Þ2 ðMw13w2MÞþ

½G=6 ðd1þd2Þ3g

whereC1;2is the coagulation rate (m33s31),Kthe attach- ment probability,Gthe shear rate (s31) andN1,N2(m33), d1,d2 (m) and w1,w2 (m s31) the particle concentrations, diameters and sinking velocities, respectively, in the two size classes. From initial experimental conditions we set G= 0.2s31;N1andN25U106cells m33and 10 aggregates m33, respectively ;d1 andd2 0.00001 and 0.002m, respec- tively ; and estimatedw1andw2from size as 0.5 and 50 m day31 according to Alldredge and Gotschalk[21].

Results of this simple model suggest that the observed decrease in the concentration of dispersed cells could be explained by aggregation due to scavenging of individual cells by large aggregates. The stickiness coe⁄cient neces- sary to explain observations assuming all cells lost from SSW aggregated was K= 0.004. This is considered a low value in the presence of high TEP concentrations [35,36]

and clearly reasonable.

3.3.3. Di¡usive e¥ux

As both biogenic silica and silicic acid concentrations in aggregates remained largely unchanged (within our ability to measure these), possible aggregation and subsequent dissolution of cells in aggregates must have been balanced by loss due to di¡usion of silicic acid out of aggregates.

Alternatively, all dissolution occurred in the SSW rather than in aggregates, implying that both, the in£ux of bio- genic silica due to aggregation and the e¥ux of silicic acid out of aggregates were negligible.

Estimates of advective or di¡usive loss of solutes from marine snow di¡er widely. Because marine aggregates have porosities s0.99 (e.g. their solid volume is small), and because neither their sinking velocity nor their coag- ulation rates with smaller particles can be predicted as- suming solid spheres [37^39], it has been argued that ad- vective and di¡usive exchange of solutes between aggregates and the surrounding water may be large (e.g.

[20]). A large fraction of the non-solid volume of aggre- gates, however, is ¢lled with a polysaccharide matrix [40,41], and especially diatom aggregates contain large amounts of mucus, which consists of TEP [42,43]. The volume fraction of aggregates, which is ¢lled by such a mucus matrix, and its impact on the e¥ux of solutes are largely unknown. Possibly this polymer network of the mucus matrix traps the interstitial water, thus e¡ectively reducing advective exchange of solutes between aggregates and the SSW. Both, the fractal nature of aggregates [23]

and this mucus matrix of aggregates could physically re- duce the e¥ux of solutes out of aggregates.

Experimental studies have shown that total oxygen ex- change between sinking aggregates and the surrounding water could largely be explained by di¡usion as the sole mass transfer process within aggregates similar to those of

(8)

the present study (Ploug et al., 2002). Aggregate volumes are in the order ofWl. Direct measurements of concentra- tions and, hence, di¡usion coe⁄cients of solutes within aggregates are therefore di⁄cult. An estimate based on point measurements of the concentration gradients of oxy- gen using oxygen microsensors within and around aggre- gates yielded a di¡usion coe⁄cient of oxygen within ag- gregates near that expected in stagnant seawater[9]. Other experiments adding tetrazolium salts to SSW, however, suggested that di¡usion rates were greatly reduced com- pared to di¡usion in stagnant seawater as the reaction with tetrazolium salts within aggregates appeared slow [7]. A di¡usion coe⁄cient for silicic acid out of natural marine snow, estimated to be at least 20^200 times less than that predicted by molecular di¡usion in seawater, was derived from a budget of the silica cycle within ag- gregates, but no time series data were available[10]. More advanced tools like nuclear magnetic resonance imaging and di¡usivity microsensors with high spatial resolution as used in sediments have not yet been used for more direct measurements of di¡usivity in diatom aggregates and marine snow[44,45].

3.4. Evaluation of the experiment on the dissolution of aggregated and dispersed diatom frustules

The experiment was conducted to look at dissolution of diatom frustules while they settle to depth. Incubation conditions mimicked sinking in the dark, deep ocean. Bot- tles were transparent for gas allowing oxygen and carbon dioxide to remain at reasonable levels. Conditions inside bottles did not deteriorate (in contrast of batch cultures at senescence). Although wall e¡ects can’t be excluded, none were observed and bacterial concentration remained real- istic, even low. Live phytoplankton cells were still ob- served after 42days of the experiment. The exclusion of grazers was a prerequisite for comparing dissolution rate of aggregated and dispersed cells. Dissolution rates of grazed frustules need to be assessed for evaluation of silica dissolution in the ocean, but were beyond the scope of this paper.

Although results appear to show that dissolution of ag- gregated diatom frustules was appreciably slower than that of dispersed cells, aggregation dynamics were so com- plex that the possibility that aggregation compensated for loss of aggregated cells due to dissolution could not be excluded. The exchange rate of silicic acid between pore water and the SSW is also too uncertain to constrain £ux of silica between aggregates and SSW. Assuming a di¡u- sion coe⁄cient of silicic acid within aggregates to be sim- ilar to molecular di¡usion in stagnant seawater, the aggre- gation rate and the subsequent e¥ux of silicic acid out of aggregates could have been high enough to generate ob- served patterns, even if dissolution rates in aggregates were higher than those of freely suspended cells. In other words, although the enrichment of cells in aggregates at the con-

clusion of the experiment suggests a higher dissolution rate for freely suspended cells, the data of this experiment can not exclude the possibility that dissolution rate of dis- persed cells was similar or even higher than that of aggre- gated cells. In a future experiment aggregated and dis- persed cells will have to be separated a priori.

3.5. Signi¢cance for ¢eld observations

Observations suggest that most of the dissolution of biogenic silica is con¢ned to a few hundred meters from the sea surface. The observed decrease in £ux of biogenic silica to depth is comparatively low[13,46^48]. High dis- solution rates of diatom frustules observed in the upper layer of the ocean have been explained by bacteria-medi- ated dissolution of unaggregated cells [4,5]. Within the euphotic zone, average dissolution rates of aggregated cells are similar [10], as high primary production [8]and high uptake rates of silicic acid by some of the aggregated cells are compensated by high dissolution rates of others [10].

We had hypothesized that low dissolution rates within aggregates could explain reduced dissolution rates below the mixed layer, because diatoms reach greater depth only via aggregates or within fecal pellets. Alternatively, the decline in silica dissolution rates with depth may be ex- plained by aggregation and subsequent rapid sedimenta- tion of cells. Changes in the average sinking velocity of diatoms due to their inclusion in fast sinking aggregates rather than di¡erences in dissolution rates would also re- sult in decreased recycling of silica at depth.

Acknowledgements

We thank Rolf Peinert, Peter Fritsche, Ursula Jung- hans, Kerstin Nachtigall, Dorothee Adams, Jenny Dann- heim and Gabi Donner for help with the sampling and Geo¡ Evans, Mark Brzezinski, Kumiko Azetsu-Scott, Bruce Logan and Dieter Wolf-Gladrow for discussions.

This research was funded by the Deutsche Forschungsge- meinschaft and the Alexander von Humboldt Foundation.

References

[1] Cappellen, P.V. and Qui, L. (1997) Biogenic silica dissolution in sedi- ments of the Southern Ocean : I Solubility. Deep-Sea Res. II 44, 1109^1128.

[2] Treguer, P. et al. (1995) The Silica balance in the world ocean : A reestimate. Science 268, 375^379.

[3] Hecky, R.E. et al. (1973) The amino acid and sugar composition of diatom cell walls. Mar. Biol. 19, 323^331.

[4] Patrick, S. and Holding, A.J. (1985) The e¡ect of bacteria on the solubilization of silica in diatom frustules. J. Appl. Bacteriol. 59, 7^

16.

[5] Bidle, K.D. and Azam, F. (1999) Accelerated dissolution of diatom silica by marine bacterial assemblages. Nature (London) 397, 508^

512.

(9)

[6] Alldredge, A.L. and Gotschalk, C.C. (1989) Direct observations of the mass £occulation of diatom blooms : Characteristics, settling ve- locities and formation of diatom aggregates. Deep-Sea Res. 36, 159^

171.

[7] Shanks, A.L. and Reeder, M.L. (1993) Reducing microzones and sul¢de production in marine snow. Mar. Ecol. Prog. Ser. 96, 43^

47.

[8] Gotschalk, C.C. and Alldredge, A.L. (1989) Enhanced primary pro- duction and nutrient regeneration within aggregated marine diatoms.

Mar. Biol. 103, 119^129.

[9] Ploug, H. et al. (1997) Anoxic aggregates ^ an ephemeral phenome- non in the ocean? Aquat. Microb. Ecol. 13, 285^294.

[10] Brzezinski, M.A., O’Bryan, L.M. and Alldredge, A.L. (1997) Silica cycling within marine snow. Limnol. Oceanogr. 42, 1706^1713.

[11] Bidle, K.D. and Azam, F. (2001) Bacterial control of silicon regen- eration from diatom detritus: Signi¢cance of bacterial ectohydrolases and species identity. Limnol. Oceanogr. 46, 1606^1623.

[12] Schrader, H.J. (1972) Anlo«sung und Konservierung von Diatomeen- schalen beim Absinken am Beispiel des Landsort-Tiefs in der Ostsee.

Nova Hedwigia 39, 191^216.

[13] Takahashi, K. (1986) Seasonal £uxes of pelagic diatoms in the sub- arctic paci¢c 1982^1983. Deep-Sea Res. I 33, 1225^1251.

[14] Guillard, R.R.L. (1975) Culture of phytoplankton for feeding marine invertebrates. In : Culture of Marine Invertebrate Animals (Smith, W.L. and Chanley, M.H., Eds.), pp. 108^132. Plenum Press, New York.

[15] Shanks, A.L. and Edmondson, E.W. (1989) Laboratory-made arti¢- cial marine snow : A biological model of the real thing. Mar. Biol.

101, 463^470.

[16] Korole¡, F. (1983) Determination of nutrients. In: Methods of Sea- water Analysis (Grassho¡, K., Erhardt, M. and Kremling, K., Eds.), pp. 125^187. Verlag Chemie, Weinheim.

[17] Parsons, T.R., Maita, Y. and Lalli, C.M. (1984) A Manual for Chem- ical and Biological Methods for Seawater Analysis. Pergamon Press, New York.

[18] Passow, U. and Alldredge, A.L. (1995) A dye-binding assay for the spectrophotometric measurement of transparent exopolymer particles (TEP). Limnol. Oceanogr. 40, 1326^1335.

[19] Passow, U. and Alldredge, A.L. (1994) Distribution, size, and bacte- rial colonization of transparent exopolymer particles (TEP) in the ocean. Mar. Ecol. Prog. Ser. 113, 185^198.

[20] Logan, B.E. and Alldredge, A.L. (1989) Potential for increased nu- trient uptake by £occulating diatoms. Mar. Biol. 101, 443^450.

[21] Alldredge, A.L. and Gotschalk, C. (1988) In situ settling behavior of marine snow. Limnol. Oceanogr. 33, 339^351.

[22] Lick, W., Huang, H. and Jepson, R. (1993) Flocculation of ¢ne- grained sediments due to di¡erential settling. J. Geophys. Res. 98, 10279^10288.

[23] Logan, B.E. and Wilkinson, B.D. (1990) Fractal geometry of marine snow and other biological aggregates. Limnol. Oceanogr. 35, 130^

135.

[24] Kilps, J.R., Logan, B.E. and Alldredge, A.L. (1994) Fractal dimen- sions of marine snow determined from image analysis of in situ pho- tographs. Deep-Sea Res. 41, 1159^1169.

[25] Jiang, Q. and Logan, B.E. (1991) Fractal dimensions of aggregates determined from steady-state size distributions. Environ. Sci. Tech- nol. 25, 2031^2038.

[26] Logan, B.E. and Kilps, J.R. (1995) Fractal dimensions of aggregates formed in di¡erent £uid mechanical environments. Water Res. 29, 443^453.

[27] Ploug, H. and Jorgensen, B.B. (1999) A net-jet £ow system for mass

transfer and microsensor studies of sinking aggregates. Mar. Ecol.

Prog. Ser. 176, 1^38.

[28] KiZrboe, T., Ploug, H. and Thygesen, U.H. (2001) Fluid motion and solute distribution around sinking aggregates. I. Small-scale £uxes and heterogeneity of nutrients in the pelagic environment. Mar.

Ecol. Prog. Ser. 211, 1^13.

[29] Logan, B.E. and Hunt, J.R. (1987) Advantages to microbes of growth in permeable aggregates in marine systems. Limnol. Ocean- ogr. 32, 1034^1048.

[30] Blank, G.S. and Sullivan, C.W. (1979) Diatom mineralization of silicic acid-III, Si(OH)4 binding and light dependent transport in Nitzschia angularis. Arch. Microbiol. 123, 157^164.

[31] Cappellen, P.V. and Qui, L. (1997) Biogenic silica dissolution of the Southern Ocean : II Kinetics. Deep-Sea Res. II 44, 1129^1149.

[32] Greenwood, J.E., Truesdale, V.W. and Rendell, A.R. (2001) Biogenic silica dissolution in seawater ^ in vitro chemical kinetics. Prog. Oce- anogr. 48, 1^23.

[33] KiZrboe, T. (2000) Colonization of marine snow aggregates by inver- tebrate zooplankton : Abundance, scaling, and possible role. Limnol.

Oceanogr. 45, 479^484.

[34] Jackson, G.A. (1990) A model of the formation of marine algal £ocks by physical coagulation processes. Deep-Sea Res. 37, 1197^1211.

[35] Dam, H.G. and Drapeau, D.T. (1995) Coagulation e⁄ciency, organ- ic-matter glues, and the dynamics of particles during a phytoplankton bloom in a mesocosm study. Deep-Sea Res. II 42, 111^123.

[36] Engel, A. (2000) The role of transparent exopolymer particles (TEP) in the increase in apparent particle stickiness (alpha) during the de- cline of a diatom bloom. J. Plankton Res. 22, 485^497.

[37] Cli¡ord, J., Li, X. and Logan, B. (1996) Settling velocities of fractal aggregates. Environ. Sci. Technol. 30, 1911^1918.

[38] Li, X. and Logan, B.E. (1997) Collision frequencies of fractal aggre- gates with small particles by di¡erential aggregation. Environ. Sci.

Technol. 31, 1229^1236.

[39] Li, X. and Logan, B.E. (1997) Collision frequencies between fractal aggregates and small particles in a turbulently sheared £uid. Environ.

Sci. Technol. 31, 1237^1242.

[40] Cowen, J.P. and Holloway, C.F. (1996) Structure and chemical anal- ysis of marine aggregates: In situ macrophotography and laser con- focal and electron microscopy. Mar. Biol. 126, 163^174.

[41] Holloway, C.F. and Cowen, J.P. (1997) Development of a scanning confocal laser microscopic technique to examine the structure and composition of marine snow. Limnol. Oceanogr. 42, 1340^1352.

[42] Alldredge, A.L., Passow, U. and Logan, B.E. (1993) The abundance and signi¢cance of a class of large, transparent organic particles in the ocean. Deep-Sea Res. I 40, 1131^1140.

[43] Passow, U., Alldredge, A.L. and Logan, B.E. (1994) The role of particulate carbohydrate exudates in the £occulation of diatom blooms. Deep-Sea Res. I 41, 335^357.

[44] Revsbech, N.P., Nielsen, L.P. and Ramsing, N.B. (1998) A novel microsensor for determination of apparent di¡usivity in sediments.

Limnol. Oceanogr. 43, 986^992.

[45] Wieland, A. et al. (2001) Fine-scale measurements of di¡usivity in a microbial mat with nuclear magnetic resonance imaging. Limnol.

Oceanogr. 46, 248^259.

[46] Honjo, S., Mahganini, S.J. and Cole, J.J. (1982) Sedimentation of biogenic matter in the deep ocean. Deep-Sea Res. 29, 609^625.

[47] da Costa DeMacho, E. (1993) Production, sedimentation and disso- lution of biogenic silica in the northern North Atlantic. In: Marine Planktology. University of Kiel, Kiel.

[48] Brzezinski, M.A. and Nelson, D.M. (1995) The annual silica cycle in the Sargasso Sea near Bermuda. Deep-Sea Res. 42, 1215^1237.

Referenzen

ÄHNLICHE DOKUMENTE

with added diatom exudates the Fe(II) concentration remained at about 30 nmol L -1 298. decreasing only very slightly

The negative trend shows a pronounced seasonality; the largest decrease occurs in autumn with magnitudes up to -4.8 %/decade (relative to the overall long-term mean

We have seen that in the five years of the Agreement, these total savings (3,330 million) were higher than those from the MSC (2,749 million), so there would have been a margin

Rezultatele mai puţin bune obţinute ca urmare a de mersurilor desfăşurate în organizațiile școlare, org anizate până în prezent cu privire la diminuarea actelor violent

• The maps show the possible decrease in the number of wet days under future climate conditions (averaged for the time period of 2036 to 2065 compared to the average of the

• Comparably low values of less than 45% decrease of the frequency of cold days are projected for large parts of Central Asia, higher decrease is projected for the southern parts

Changes in fetal aortic blood pressure (PAorta) in mm Hg and the blood flow in one umbilical artery and vein each (QuA and Quv) in ml/min with maternal hypoxia (experiments Hl—H7)

Das „saldenmechanische Modell“ von Fritz Helmedag und die Empirie Helmedag hat im „Wirtschaftsdienst“ unter der Rubrik „Wissenschaft für die Praxis“ eine Variante